Skip to main content Accessibility help
×
Hostname: page-component-7bb8b95d7b-cx56b Total loading time: 0 Render date: 2024-09-12T20:06:35.187Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  18 January 2010

Christian de Duve
Affiliation:
Rockefeller University, New York
Get access
Type
Chapter
Information
Singularities
Landmarks on the Pathways of Life
, pp. 241 - 250
Publisher: Cambridge University Press
Print publication year: 2005

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

, A. Akhmanova, , F. Voncken, , T. van Halen, , A. van Hoek, , B. Boxma, , G. Vogels, , M. Veenhuis, and , H. H. P. Hackstein (1998). A hydrogenosome with a genome. Nature, 396, 527–528CrossRefGoogle Scholar
, A. D. Anbar and , A. H. Knoll (2002). Proterozoic ocean chemistry and evolution: A bioinorganic bridge. Science, 297, 1137–1142CrossRefGoogle Scholar
, J. O. Andersson, , A. M. Sjögren, , L. A. M. Davis, , T. M. Embley, and , A. J. Roger (2003). Phylogenetic analyses of diplomonad genes reveal frequent lateral gene transfers affecting eukaryotes. Curr. Biol., 13, 94–104CrossRefGoogle Scholar
, L. Aravind, , R. L. Tatusov, , Y. I. Wolf, , D. R. Walker, and , E. V. Koonin (1998). Evidence for massive gene exchange between archaeal and bacterial hyperthermophiles. Trends Genet., 14, 442–444CrossRefGoogle Scholar
, G. L. Arnold, , A. D. Anbar, , J. Barling, and , T. W. Lyons (2004). Molybdenum isotope evidence for widespread anoxia in mid-proterozoic oceans. Science, 304, 87–90CrossRefGoogle Scholar
G. Arrhenius, B. Gedulin, and S. Mojzsis (1993). Phosphate in models for chemical evolution. In Chemical Evolution and Origin of Life (, C. Ponnamperuma and , J. Chela-Flores, eds.), 25–50. Hampton VA: A. Deepak PublishersGoogle Scholar
, M. Balter (2000). Evolution on life's fringes. Science, 289, 1866–1867CrossRefGoogle Scholar
M. Baltscheffsky and H. Baltscheffsky (1992). Inorganic pyrophosphate and inorganic pyrophosphatases. In Molecular Mechanisms in Bioenergetics (, L. Ernster, ed.), 331–348. Amsterdam: ElsevierGoogle Scholar
, A. Bar-Nun, , E. Kochavi, and , S. Bar-Nun (1994). Assemblies of free amino acids as possible prebiotic catalysts. J. Mol. Evol., 39, 116–122Google Scholar
, A. D. Baughn and , M. H. Malamy (2004). The strict anaerobe Bacteroides fragilis grows in and benefits from nanomolar concentrations of oxygen. Nature, 427, 441–444CrossRefGoogle Scholar
, A. Bekker, , H. D. Holland, , P.-L. Wang, , D. Rumble III, , H. J. Stein, , J. L. Hannah, , L. L. Coetzee, and , N. J. Beukes (2004). Dating the rise of atmospheric oxygen. Nature, 427, 117–120CrossRefGoogle Scholar
, M. P. Bernstein, , J. P. Dworkin, , S. A. Sandford, , G. W. Cooper, and , L. J. Allamandola (2002). Racemic amino acids from the ultraviolet photolysis of interstellar ice analogues. Nature, 416, 401–403CrossRefGoogle Scholar
, O. Botta and , J. L. Bada (2002). Extraterrestrial organic compounds in meteorites. Surv. Geophys., 23, 411–467CrossRefGoogle Scholar
, O. Botta, , D. P. Glavin, , G. Kminek, and , J. L. Bada (2002). Relative amino acid concentrations as a signature for parent body processes of carbonaceous chondrites. Orig. Life Evol. Biosph., 32, 143–164CrossRefGoogle Scholar
, Y. Boucher, , M. Kamekura, and , W. F. Doolittle (2004). Origins and evolution of isoprenoid lipid biosynthesis in archaea. Mol. Microbiol., 52 (No. 2), 515–527CrossRefGoogle Scholar
A. Brack (2003). La chimie de l'origine de la vie. In Les Traces du Vivant (, M. Gargaud, , D. Despois, , J.-P. Parisot, and , J. Reisse, eds.), 61–81. Pessac: Presses Universitaires de BordeauxGoogle Scholar
, M. D. Brasier, , O. R. Green, , A. P. Jephcoat, , A. K. Kleppe, , M. J. Van Kranendonk, , J. F. Lindsay, , A. Steele, and , N. V. Grassineau (2002). Questioning the evidence for Earth's oldest fossils. Nature, 416, 76–81CrossRefGoogle Scholar
, J. J. Brocks, , G. A. Logan, , R. Buick, and , R. E. Summons (1999). Archaean molecular fossils and the early rise of eukaryotes. Science, 285, 1033–1036CrossRefGoogle Scholar
, P. Brown, , T. Sutikna, , M. J. Morwood, , R. P. Soejono, , E. Wayhu Saptomo, and , R. Awe Due (2004). A new small-bodied hominin from the late Pleistocene of Flores, Indonesia. Nature, 431, 1055–1061CrossRefGoogle Scholar
, R. Cammack (1983). Evolution and diversity in the iron–sulfur proteins. Chem. Scr., 21, 87–95Google Scholar
, D. E. Cane (1997). Polyketide and nonribosomal polypeptide biosynthesis. Chem. Rev., 97, 2463–2706. (includes 13 papers on the topic.)CrossRefGoogle Scholar
, D. E. Canfield (1998). A new model for Proterozoic ocean chemistry. Nature, 396, 450–453CrossRefGoogle Scholar
, L. H. Caporale (2003). Foresight in genome evolution. Am. Sci., 91, 234–241CrossRefGoogle Scholar
, D. Caramelli, , C. Lalueza-Fox, , C. Vernes, , M. Lari, , A. Casoli, , F. Mallegni, , B. Chiarelli, , I. Dupanloup, , J. Bertranpetit, , G. Barbujani, and , G. Bertorelle (2003). Evidence for a genetic discontinuity between Neandertals and 24,000-year-old anatomically modern Europeans. Proc. Nat. Acad. Sci. U.S.A., 100, 6593–6597CrossRefGoogle Scholar
, S. B. Carroll (2003). Genetics and the making of Homo sapiens. Nature, 422, 849–857CrossRefGoogle Scholar
, J. Castresana and , M. Saraste (1995). Evolution of energetic metabolism: The respiration early hypothesis. Trends Biol. Sci., 20, 443–448CrossRefGoogle Scholar
, T. Cavalier-Smith (1975). The origin of nuclei and eukaryotic cells. Nature, 256, 463–468CrossRefGoogle Scholar
, T. Cavalier-Smith (2002). The phagotrophic origin of eukaryotes and phylogenetic classification of Protozoa. Int. J. Syst. Evol. Microbiol., 52, 297–354CrossRefGoogle Scholar
, T. R. Cech (1986). RNA as an enzyme. Sci. Am., 255 (No. 5), 64–75CrossRefGoogle Scholar
, A. Chakrabarti, , R. R. Breaker, , G. F. Joyce, and , D. W. Deamer (1994). Production of RNA by a polymerase protein encapsulated within phospholipid vesicles. J. Mol. Evol., 39, 555–559CrossRefGoogle Scholar
, S. Conway Morris (1998). The Crucible of Creation. Oxford: Oxford University PressGoogle Scholar
, S. Conway Morris (2003). Life's Solution. Cambridge: Cambridge University PressGoogle Scholar
, J. R. Cronin and , S. Pizzarello (1997). Enantiomeric excesses in meteoritic amino acids. Science, 275, 951–955CrossRefGoogle Scholar
, C. Cunchillos and , G. Lecointre (2002). Early steps of metabolism evolution inferred by cladistic analysis of amino acid catabolic pathways. C. R. Biol., 325, 119–129CrossRefGoogle Scholar
, C. Cunchillos and , G. Lecointre (2005). Integrating the universal metabolism into a phylogenetic analysis. Mol. Biol. Evol., 22, 1–11Google Scholar
D. W. Deamer (1998). Membrane compartments in prebiotic evolution. In The Molecular Origins of Life (, A. Brack, ed.), 189–205. Cambridge: Cambridge University PressGoogle Scholar
C. de Duve (1963). The lysosome concept. In Ciba Foundation Symposium on Lysosomes (, A. V. S. de Reuck and , M. P. Cameron, eds.), 1–31. London: ChurchillGoogle Scholar
, C. de Duve (1969). Evolution of the peroxisome. Ann. N.Y. Acad. Sci., 168, 369–381CrossRefGoogle Scholar
, C. de Duve (1984). A Guided Tour of the Living Cell. New York: Scientific American BooksGoogle Scholar
, C. de Duve (1988). The second genetic code. Nature, 333, 117–118CrossRefGoogle Scholar
, C. de Duve (1991). Blueprint for a Cell. Burlington, NC: Neil Patterson Publishers, Carolina Biological Supply CompanyGoogle Scholar
, C. de Duve (1993). The RNA world: Before and after?Gene, 135, 29–31CrossRefGoogle Scholar
, C. de Duve (1995). Vital Dust. New York: BasicBooksGoogle Scholar
, C. de Duve (1996). The birth of complex cells. Sci. Am., 274 (No. 4), 38–45Google Scholar
C. de Duve (1998). Clues from present-day biology: The thioester world. In The Molecular Origins of Life (, A. Brack, ed.), 219–236. Cambridge: Cambridge University PressGoogle Scholar
C. de Duve (2001). The origin of life: Energy. In Frontiers of Life. (, D. Baltimore, , R. Dulbecco, , F. Jacob, and , R. Levi-Montalcini, eds.), Vol. I, 153–168. San Diego, CA: Academic PressGoogle Scholar
, C. de Duve (2002). Life Evolving. New York: Oxford University PressGoogle Scholar
, C. de Duve (2003). A research proposal on the origin of life. Orig. Life Evol. Biosph., 33, 1–16CrossRefGoogle Scholar
, C. de Duve (2005). The onset of selection. Nature, 433, 581–582CrossRefGoogle Scholar
, C. de Duve and , R. Wattiaux (1966). Functions of lysosomes. Annu. Rev. Physiol., 28, 435–492CrossRefGoogle Scholar
, F. M. Devienne, , C. Barnabé, , M. Couderc, and , G. Ourisson (1998). Synthesis of biological compounds in quasi-interstellar conditions. C. R. Acad. Sci. Paris, Sér. IIc, 1, 435–439Google Scholar
, F. M. Devienne, , C. Barnabé, and , G. Ourisson (2002). Synthesis of further biological compounds in interstellar-like conditions. C. R. Acad. Sci. Paris, Chimie/Chemistry, 5, 651–653Google Scholar
, R. E. Dickerson and , I. Geis (1969). The Structure and Action of Proteins. Menlo Park, CA: Benjamin/Cummings Publishing CompanyGoogle Scholar
, M. Di Giulio (2003). The early phases of genetic code origin: Conjectures on the evolution of coded catalysis. Orig. Life Evol. Biosph., 33, 479–489CrossRefGoogle Scholar
, W. F. Doolittle (1999). Phylogenetic classification and the universal tree. Science, 284, 2124–2128CrossRefGoogle Scholar
, W. F. Doolittle (2000). The nature of the universal ancestor and the evolution of the proteome. Curr. Opin. Struct. Biol., 10, 355–358CrossRefGoogle Scholar
, S. D. Dyall, , M. T. Brown, and , P. J. Johnson (2004a). Ancient invasions: From endosymbionts to organelles. Science, 304, 253–257CrossRefGoogle Scholar
, S. D. Dyall, , W. Yan, , M. G. Delgadillo-Correa, , A. Lunceford, , J. A. Loo, , C. F. Clarke, and , P. J. Johnson (2004b). Non-mitochondrial complex I proteins in a Trichomonas hydrogenosomal oxidoreductase complex. Nature, 431, 1103–1107CrossRefGoogle Scholar
, P. Ehrenfreund, , W. Irvine, , L. Becker, , J. Blank, , J. R. Brucato, , L. Colangeli, , S. Derenne, , D. Despois, , A. Dutrey, , H. Fraaije, , A. Lazcano, , T. Owen, and , F. Robert, an International Space Science Institute ISSI Team (2002). Astrophysical and astrochemical insights into the origin of life. Rep. Prog. Phys., 65, 1427–1487CrossRefGoogle Scholar
, M. Eigen and , P. Schuster (1977). The hypercycle: A principle of self-organization. Part A: Emergence of the hypercycle. Naturwissenschaften, 64, 541–565CrossRefGoogle Scholar
, M. Eigen and , R. Winkler-Oswatitsch (1981). Transfer-RNA, an early gene. Naturwissenschaften, 68, 282–292CrossRefGoogle Scholar
, A. Eschenmoser (1999). Chemical etiology of nucleic acid structure. Science, 284, 2118–2124CrossRefGoogle Scholar
, P. Forterre (1995). Thermoreduction, a hypothesis for the origin of prokaryotes. C. R. Acad. Sci. III, 318, 415–422Google Scholar
, P. Forterre (1999). Where is the root of the universal tree of life?BioEssays, 21, 871–8793.0.CO;2-Q>CrossRefGoogle Scholar
, P. Forterre, , C. Bouthier de la Tour, , H. Philippe, and , M. Duguet (2000). Reverse gyrase from hyperthermophiles: Probable transfer of a thermoadaptation trait from Archaea to Bacteria. Trends Genet., 16, 152–154CrossRefGoogle Scholar
, S. W. Fox (1988). The Emergence of Life. New York: BasicBooksGoogle Scholar
, S. J. Freeland, , T. Wu, and , N. Keulmann (2003). The case for an error minimizing standard genetic code. Orig. Life Evol. Biosph., 33, 457–477CrossRefGoogle Scholar
, N. Fujii (2002). d-Amino acids in living higher organisms. Orig. Life Evol. Biosph., 32, 103–127CrossRefGoogle Scholar
, H. Furnes, , N. R. Banerjee, , K. Muehlenbachs, , H. Staudigel, and , M. de Wit (2004). Early life recorded in archean pillow lavas. Science, 304, 578–581CrossRefGoogle Scholar
, N. Galtier, , N. Tourasse, and , M. Gouy (1999). A nonhyperthermophilic common ancestor to extant life forms. Science, 283, 220–221CrossRefGoogle Scholar
, J. M. Garcia-Ruiz, , S. T. Hyde, , A. M. Carnerup, , A. G. Christy, , M. J. Van Kranendonk, and , N. J. Welham (2003). Self-assembled silica–carbonate structures and detection of ancient microfossils. Science, 302, 1194–1197CrossRefGoogle Scholar
B. Gedulin and G. Arrhenius (1994). Sources and geochemical evolution of RNA precursor molecules – the role of phosphate. In Early Life on Earth, Nobel Symposium 84 (, S. Bengtson, ed.), 91–110. New York: Columbia University PressGoogle Scholar
, W. Gilbert (1986). The RNA world. Nature, 319, 618CrossRefGoogle Scholar
, W. Gilbert, , M. Marchionni, and , G. Mcknight (1986). On the antiquity of introns. Cell, 46, 151–154CrossRefGoogle Scholar
, N. Glansdorff (2000). About the last common ancestor, the universal life-tree, and lateral gene transfer: A reappraisal. Mol. Microbiol., 38 (No. 2), 177–185CrossRefGoogle Scholar
, M. Gogarten-Boekels, , E. Hilario, and , J. P. Gogarten (1995). The effects of heavy meteorite bombardment on the early evolution – the emergence of the three domains of life. Orig. Life Evol. Biosph., 25, 251–264CrossRefGoogle Scholar
, S. Gribaldo and , H. Philippe (2002). Ancient phylogenetic relationships. Theor. Pop. Biol., 61, 391–408CrossRefGoogle Scholar
, R. S. Gupta, , K. Aitken, , M. Falah, and , B. Singh (1994). Cloning of Giardia lamblia heat shock protein HSP70 homologs: Implications regarding origin of eukaryotic cells and endoplasmic reticulum. Proc. Nat. Acad. Sci. U.S.A., 91, 2895–2899CrossRefGoogle Scholar
, J. H. P. Hackstein, , A. Akhmanova, , F. Voncken, , A. van Hoek, , T. van Alen, , B. Boxma, , S. Y. Moon-van der Staay, , G. van der Staay, , J. Leunissen, , M. Huynen, , J. Rosenberg, and , M. Veenhuis (2001). Hydrogenosomes: Convergent adaptations of mitochondria to anaerobic environments. Zoology, 104, 290–302CrossRefGoogle Scholar
, M. M. Hanczyc, , S. M. Fujikawa, and , J. W. Szostak (2003). Experimental models of primitive cellular compartments: Encapsulation, growth, and division. Science, 302, 618–622CrossRefGoogle Scholar
, H. Hartman (1984). The origin of the eukaryotic cell. Speculations Sci. Technol., 7 (No. 2), 77–81Google Scholar
, H. Hartman and , A. Fedorov (2002). The origin of the eukaryotic cell: A genomic investigation. Proc. Nat. Acad. Sci. U.S.A., 99, 1420–1425CrossRefGoogle Scholar
, T. Horiike, , K. Hamada, , S. Kanaya, and , T. Shinozawa (2001). Origin of eukaryotic cell nuclei by symbiosis of Archaea in Bacteria is revealed by homology-hit analysis. Nature Cell Biol., 3, 210–214CrossRefGoogle Scholar
, D. S. Horner, , P. G. Foster, and , T. M. Embley (2000). Iron hydrogenases and the evolution of anaerobic eukaryotes. Mol. Biol. Evol., 17 (No. 11), 1695–1709CrossRefGoogle Scholar
, C. Huber and , G. Wächtershäuser (1998). Peptides by activation of amino acids by CO on (Ni, Fe)S surfaces: Implications for the origin of life. Science, 281, 670–672CrossRefGoogle Scholar
, E. Imai, , H. Honda, , K. Hatori, , A. Brack, and , K. Matsuno (1999). Elongation of oligopeptides in a simulated submarine hydrothermal system. Science, 283, 831–833CrossRefGoogle Scholar
, W. K. Johnston, , P. J. Unrau, , M. S. Lawrence, , M. E. Glasner, and , D. P. Bartel (2001). RNA-catalyzed RNA polymerization: Accurate and general RNA-templated primer extension. Science, 292, 1319–1325CrossRefGoogle Scholar
, A. Jorissen and , C. Cerf (2002). Asymmetric photoreactions as the origin of biomolecular homochirality: A critical review. Orig. Life Evol. Biosph., 32, 129–142CrossRefGoogle Scholar
, G. F. Joyce (2002). The antiquity of RNA-based evolution. Nature, 418, 214–221CrossRefGoogle Scholar
G. F. Joyce and L. E. Orgel (1993). Prospects for the understanding of the origin of the RNA world. In The RNA World (, R. F. Gesteland and , J. F. Atkins, eds.), 1–25. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory PressGoogle Scholar
, O. Kandler (1994a). Cell wall biochemistry in Archaea and its phylogenetic implications. J. Biol. Phys., 20, 165–169Google Scholar
O. Kandler (1994b). The early diversification of life. In Early Life on Earth, Nobel Symposium 84 (, S. Bengtson, ed.), 152–160. New York: Columbia University PressGoogle Scholar
, K. Kashefi and , D. R. Lovley (2003). Extending the upper temperature limit for life. Science, 301, 934CrossRefGoogle Scholar
M. Kates (1992). Archaebacterial lipids: Structure, biosynthesis and function. In The Archaebacteria: Biochemistry and Biotechnology (, M. J. Danson, , D. W. Hough, and , G. G. Lunt, eds.), 51–72. Biochem. Soc. Symp. 58. London: Portland PressGoogle Scholar
, A. D. Keefe and , S. L. Miller (1995). Are polyphosphates or phosphate esters prebiotic reagents?J. Mol. Evol., 41, 693–702Google Scholar
, A. D. Keefe, , G. L. Newton, and , S. L. Miller (1995). A possible prebiotic synthesis of pantetheine, a precursor of coenzyme A. Nature, 373, 683–685CrossRefGoogle Scholar
, A. H. Knoll (2003). Life on a Young Planet. Princeton, NJ: Princeton University PressGoogle Scholar
, V. Kolb, , S. Zhang, , Y. Xu, and , G. Arrhenius (1997). Mineral-induced phosphorylation of glycolate ion – a metaphor in chemical evolution. Orig. Life Evol. Biosph., 27, 485–503CrossRefGoogle Scholar
, E. V. Koonin (2003). Comparative genomics, minimal gene-sets and the last universal common ancestor. Nature Rev. Microbiol., 1, 127–136CrossRefGoogle Scholar
, A. Kornberg, , N. N. Rao, and , D. Ault-Riché (1999). Inorganic polyphosphate: A molecule of many functions. Annu. Rev. Biochem., 68, 89–125CrossRefGoogle Scholar
, I. S. Kulaev (1979). The Biochemistry of Inorganic Polyphosphates. New York: WileyGoogle Scholar
, J. A. Lake, , R. Jain, and , M. C. Rivera (1999). Mix and match in the tree of life. Science, 283, 2027–2028CrossRefGoogle Scholar
, N. Lane (2002). Oxygen. Oxford: Oxford University PressGoogle Scholar
, A. Lazcano (2003). Just how pregnant is the universe?Science, 299, 347–348CrossRefGoogle Scholar
, L. Leman, , L. Orgel, and , M. Reza Ghadiri (2004). Carbonyl sulfide-mediated prebiotic formation of peptides. Science, 306, 283–286CrossRefGoogle Scholar
, R. Lewin (1996). Patterns in Evolution. New York: Scientific American BooksGoogle Scholar
, M. R. Lindsay, , R. I. Webb, , M. Strous, , M. S. Jetten, , M. K. Butler, , R. J. Forde, and , J. A. Fuerst (2001). Cell compartmentalisation in planctomycetes: Novel types of structural organisation for the bacterial cell. Arch. Microbiol., 175 (No. 6), 413–429CrossRefGoogle Scholar
P. L. Luisi (2002). Some open questions about the origin of life. In Fundamentals of Life (, G. Palyi, , C. Zucchi, and , L. Caglioti, eds.), 289–301. Paris: ElsevierGoogle Scholar
, D. A. Mac Donaill (2003). Why nature chose A, C, G, and U/T: An error-coding perspective of nucleotide alphabet composition. Orig. Life Evol. Biosph., 33, 433–455CrossRefGoogle Scholar
, L. Margulis (1981). Symbiosis in Cell Evolution. San Francisco: W. H. Freeman & CoGoogle Scholar
, L. Margulis (1996). Archaeal–eubacterial mergers in the origin of Eukarya: Phylogenetic classification of life. Proc. Nat. Acad. Sci. U.S.A., 93, 1071–1076CrossRefGoogle Scholar
, L. Margulis and , D. Sagan (1986). Micro-cosmos. New York: Summit BooksGoogle Scholar
, W. Martin (1999). Mosaic bacterial chromosomes: A challenge en route to a tree of genomes. BioEssays, 21, 99–1043.0.CO;2-B>CrossRefGoogle Scholar
, W. Martin and , M. Müller (1998). The hydrogen hypothesis for the first eukaryote. Nature, 392, 37–41CrossRefGoogle Scholar
, W. Martin, , C. Rotte, , M. Hoffmeister, , U. Theissen, , G. Gelius-Dietrich, , S. Ahr, and , K. Henze (2003). Early cell evolution, eukaryotes, anoxia, sulfide, oxygen, fungi first (?), and a tree of genomes revisited. Life, 55 (No. 4–5), 193–204Google Scholar
, W. Martin and , M. J. Russell (2003). On the origins of cells: A hypothesis for the evolutionary transitions from abiotic geochemistry to chemoautotrophic prokaryotes, and from prokaryotes to nucleated cells. Phil. Trans. R. Soc. London B, 358, 59–85CrossRefGoogle Scholar
, S. L. Miller (1953). A production of amino acids under possible primitive earth conditions. Science, 117, 528–529CrossRefGoogle Scholar
, K. Miller (1999). Finding Darwin's God. New York: HarperCollinsGoogle Scholar
, M. Mirazon Lahr and , R. Foley (2004). Human evolution writ small. Nature, 431, 1043–1044CrossRefGoogle Scholar
, P. A. Monnard, , C. L. Apel, , A. Kanavarioti, and , D. W. Deamer (2004). Influence of ionic inorganic solutes on self-assembly and polymerization processes related to early forms of life: Implications for a prebiotic aqueous medium. Astrobiology, 2 (No. 2), 139–152Google Scholar
, D. Moreira and , P. Lopez-Garcia (1998). Symbiosis between methanogenic archaea and δ-proteobacteria as the origin of eukaryotes: The syntrophic hypothesis. J. Mol. Evol., 47, 517–530CrossRefGoogle Scholar
, H. J. Morowitz (1999). A theory of biochemical organization, metabolic pathways, and evolution. Complexity, 4 (No. 6), 39–533.0.CO;2-2>CrossRefGoogle Scholar
, M. J. Morwood, , R. P. Soejono, , R. G. Roberts, , T. Sutikna, , C. S. M. Turney, , K. E. Westaway, , W. J. Rink, , J.-x. Zhao, , G. D. van den Bergh, , R. Awe Due, , D. R. Hobbs, , M. W. Moore, , M. I. Bird, and , L. K. Fifield (2004). Archaeology and age of a new hominin from Flores in eastern Indonesia. Nature, 431, 1087–1091CrossRefGoogle Scholar
, M. Müller (1993). The hydrogenosome. J. Gen. Microbiol., 139, 2879–2889CrossRefGoogle Scholar
, M. Müller and , W. Martin (1999). The genome of Rickettsia prowazekii and some thoughts on the origin of mitochondria and hydrogenosomes. BioEssays, 21, 377–3813.0.CO;2-W>CrossRefGoogle Scholar
, G. M. Munoz Caro, , U. J. Meierhenrich, , W. A. Schutte, , B. Barbier, , A. Arcones Segovia, , H. Rosenbauer, , W. H.-P. Thiemann, , A. Brack, and , J. M. Greenberg (2002). Amino acids from ultraviolet irradiation of interstellar ice analogues. Nature, 416, 403–406CrossRefGoogle Scholar
, K. E. Nelson, , R. A. Clayton, , S. R. Gill, , M. L. Gwinn, , R. J. Dodson, , D. H. Haft, , E. K. Hickey, , J. D. Peterson, , W. C. Nelson, , K. A. Ketchum, , L. McDonald, , T. R. Utterback, , J. A. Malek, , K. D. Linher, , M. M. Garrett, , A. M. Stewart, , M. D. Cotton, , M. S. Pratt, , C. A. Phillips, , D. Richardson, , J. Heidelberg, , G. G. Sutton, , R. D. Fleischmann, , J. A. Eisen, , O. White, , S. L. Salzberg, , H. O. Smith, , J. C. Venter, and , C. M. Fraser (1999). Evidence for lateral gene transfer between Archaea and Bacteria from genome sequence of Thermotoga maritima. Nature, 399, 323–329CrossRefGoogle Scholar
, E. Nevo (1999). Mosaic Evolution of Subterranean Mammals. Oxford: Oxford University PressGoogle Scholar
, E. Nogales, , K. H. Downing, , L. A. Amos, and , J. Löwe (1998). Tubulin and FtsZ form a distinct family of GTPases. Nature Struct. Biol., 5, 451–458CrossRefGoogle Scholar
, H. Ochman, , J. G. Lawrence, and , E. A. Groisman (2000). Lateral gene transfer and the nature of bacterial evolution. Nature, 405, 299–304CrossRefGoogle Scholar
, H. Ogasawara, , A. Yoshida, , E. Imai, , H. Honda, , K. Hatori, and , K. Matsuno (2000). Synthesizing oligomers from monomeric nucleotides in simulated hydrothermal environments. Orig. Life Evol. Biosph., 30, 519–526CrossRefGoogle Scholar
, Y. Ogata, , E. Imai, , H. Honda, , K. Hatori, and , K. Matsuno (2000). Hydrothermal circulation of sea water through hot vents and contribution of interface chemistry to prebiotic synthesis. Orig. Life Evol. Biosph., 30, 527–537CrossRefGoogle Scholar
, L. E. Orgel (2003). Some consequences of the RNA world hypothesis. Orig. Life Evol. Biosph., 33, 211–218CrossRefGoogle Scholar
S. Osawa (1995). Evolution of the Genetic Code. Oxford University Press
, G. Ourisson and , T. Nakatani (1994). The terpenoid theory of the origin of cellular life: The evolution of terpenoids to cholesterol. Chemistry and Biology, 1, 11–23CrossRefGoogle Scholar
, K. Ozawa, , A. Nemoto, , E. Imai, , H. Honda, , K. Hatori, and , K. Matsuno (2004). Phosphorylation of nucleotide molecules in hydrothermal environments. Orig. Life Evol. Biosph., 34, 465–471CrossRefGoogle Scholar
, J. D. Palmer (2003). The symbiotic birth and spread of plastids: How many times and whodunit?J. Phycol., 39, 4–11CrossRefGoogle Scholar
, S. Pitsch, , A. Eschenmoser, , B. Gedulin, , S. Hui, and , G. Arrhenius (1995). Mineral induced formation of sugar phosphates. Orig. Life Evol. Biosph., 25, 294–334Google Scholar
, S. Pizzarello and , A. L. Weber (2004). Prebiotic amino acids as asymmetric catalysts. Science, 303, 1151CrossRefGoogle Scholar
, A. Poole, , D. Jeffares, and , D. Penny (1999). Early evolution: prokaryotes, the new kids on the block. Bioessays, 21, 880–8893.0.CO;2-P>CrossRefGoogle Scholar
, D. Prangishvili (2003). Evolutionary insights from studies on viruses of hyperthermophilic archaea. Res. Microbiol., 154, 289–294CrossRefGoogle Scholar
, B. P. Prieur (2001). Étude de l'activité prébiotique potentielle de l'acide borique. C. R. Acad. Sci. Paris, Chimie/Chemistry, 4, 1–4Google Scholar
, A. Ricardo, , M. A. Carrigan, , A. N. Olcott, and , S. A. Benner (2004). Borate minerals stabilize ribose. Science, 303, 196CrossRefGoogle Scholar
, M. C. Rivera and , J. A. Lake (2004). The ring of life provides evidence for a genome fusion origin of eukaryotes. Nature, 431, 152–155CrossRefGoogle Scholar
, M. Rohmer (1999). The discovery of a mevalonate-independent pathway for isoprenoid biosynthesis in bacteria, algae, and higher plants. Nat. Prod. Rep., 16, 565–574CrossRefGoogle Scholar
, M. Rohmer, , C. Grosdemange-Billiard, , M. Seemann, and , D. Tritsch (2004). Isoprenoid biosynthesis as a novel target for antibacterial and antiparasitic drugs. Curr. Opin. Investig. Drugs, 5 (No. 2), 154–162Google Scholar
, L. Sagan (1967). On the origin of mitosing cells. J. Theor. Biol., 14, 225–274CrossRefGoogle Scholar
, J. G. Schmidt, , P. E. Nielsen, and , L. E. Orgel (1997). Enantiomeric cross-inhibition in the synthesis of oligonucleotides on a nonchiral template. J. Am. Chem. Soc., 119, 1494–1495CrossRefGoogle Scholar
, J. W. Schopf (1999). Cradle of Life. Princeton, NJ: Princeton University PressGoogle Scholar
A. W. Schwartz (1998). Origins of the RNA world. In The Molecular Origins of Life (, A. Brack, ed.), 237–254. Cambridge: Cambridge University PressGoogle Scholar
, M. Shimizu (1982). Molecular basis for the genetic code. J. Mol. Evol., 18, 297–303CrossRefGoogle Scholar
, A. Shimoyama and , R. Ogasawara (2002). Peptides and diketopiperazines in the Yamato-791198 and Murchison carbonaceous chondrites. Orig. Life Evol. Biosph., 32, 165–179CrossRefGoogle Scholar
, N. H. Sleep, , K. J. Zahnle, , J. F. Kasting, and , H. J. Morowitz (1989). Annihilation of ecosystems by large asteroid impacts on the early Earth. Nature, 342, 139–142CrossRefGoogle Scholar
M. L. Sogin, J. D. Silberman, G. Hinkle, and H. G. Morrison (1996). Problems with molecular diversity in the Eukarya. In Evolution of Microbial Life (, D. McL. Roberts, , P. Sharp, , G. Alderson, and , M. A. Collins, eds.), 167–184. Cambridge: Cambridge University PressGoogle Scholar
, S. Spiegelman (1967). An in vitro analysis of a replicating molecule. Am. Sci., 55, 221–264Google Scholar
, R. Y. Stanier (1970). Some aspects of the biology of cells and their possible evolutionary significance. Symp. Soc. Gen. Microbiol., 20, 1–38Google Scholar
, R. Sutak, , P. Dolezal, , H. L. Fiumera, , I. Hrdy, , A. Dancys, , M. Delgadillo-Correa, , P. J. Johnson, , M. Müller, and , J. Tachezy (2004). Mitochondrial-type assembly of Fe–S centers in the hydrogenosomes of the amitochondriate eukaryote Trichomonas vaginalis. Proc. Nat. Acad. Sci. U.S.A., 101, 10368–10373CrossRefGoogle Scholar
, E. Szathmary (2002). The gospel of inevitability. Nature, 419, 779–780CrossRefGoogle Scholar
, J. W. Szostak, , D. P. Bartel, and , P. L. Luisi (2001). Synthesizing life. Nature, 409, 387–390CrossRefGoogle Scholar
, K. Tamura and , P. Schimmel (2004). Chiral-selective aminoacylation of an RNA minihelix. Science, 305, 1253CrossRefGoogle Scholar
, J. Tovar, , G. Leon-Avila, , L. B. Sanchez, , R. Sutak, , J. Tachezy, , M. van der Giezen, , M. Hernandez, , M. Müller, and , J. M. Lucocq (2003). Mitochondrial remnant organelles of Giardia function in iron–sulphur protein maturation. Nature, 426, 172–176CrossRefGoogle Scholar
, F. van den Ent, , L. A. Amos, and , J. Löwe (2001). Prokaryotic origin of the actin cytoskeleton. Nature, 413, 39–44CrossRefGoogle Scholar
, M. A. van Zullen, , A. Lepland, and , G. Arrhenius (2002). Reassessing the evidence for the earliest traces of life. Nature, 418, 627–630CrossRefGoogle Scholar
, T. Vellai and , G. Vida (1999). The origin of eukaryotes: The difference between prokaryotic and eukaryotic cells. Proc. R. Soc. London B, 266, 1571–1577CrossRefGoogle Scholar
, F. Voncken, , B. Boxma, , J. Tjaden, , A. Akhmanova, , M. Huynen, , F. Verbeek, , A. G. M. Tielens, , I. Haferkamp, , H. E. Neuhaus, , G. Vogels, , M. Veenhuis, and , J. H. P. Hackstein (2002). Multiple origins of hydrogenosomes: Functional and phylogenetic evidence from the ADP/ATP carrier of the anaerobic chytrid Neocallimastix sp. Mol. Microbiol., 44 (No. 6), 1441–1454CrossRefGoogle Scholar
, C. D. von Dohlen, , S. Kohler, , S. T. Alsop, and , W. R. McManus (2001). Mealybug beta-proteobacterial endosymbionts contain gamma-proteobacterial symbionts. Nature, 412, 433–436CrossRefGoogle Scholar
G. Wächtershäuser (1998). Origin of life in an iron–sulfur world. In The Molecular Origins of Life (, A. Brack, ed.), 206–218. Cambridge: Cambridge University PressGoogle Scholar
, G. Wächtershäuser (2003). From pre-cells to Eukarya – a tale of two lipids. Mol. Microbiol., 47 (No. 1), 13–22Google Scholar
, C. T. Walsh (2004). Polyketide and nonribosomal peptide antibiotics: Modularity and versatility. Science, 303, 1805–1810CrossRefGoogle Scholar
, J. Washington (2000). The possible role of volcanic aquifers in prebiologic genesis of organic compounds and RNA. Orig. Life Evol. Biosph., 30, 53–79CrossRefGoogle Scholar
, A. L. Weber (2001). The sugar model: Catalysis by amines and amino acid products. Orig. Life Evol. Biosph., 31, 71–86CrossRefGoogle Scholar
, F. H. Westheimer (1987). Why nature chose phosphates. Science, 235, 1173–1178CrossRefGoogle Scholar
, T. D. White, , B. Asfaw, , D. DeGusta, , H. Gilbert, , G. D. Richards, , G. Suwa, and , F. C. Howell (2003). Pleistocene Homo sapiens from Middle Awash, Ethiopea. Nature, 423, 742–747CrossRefGoogle Scholar
, J. Whitfield (2004). Born in a watery commune. Nature, 427, 674–676CrossRefGoogle Scholar
, C. R. Woese (1987). Bacterial evolution. Microbiol. Rev., 51, 221–271Google Scholar
, C. R. Woese (1998). The universal ancestor. Proc. Nat. Acad. Sci. U.S.A., 95, 6854–6859CrossRefGoogle Scholar
, C. R. Woese (2000). Interpreting the universal phylogenetic tree. Proc. Nat. Acad. Sci. U.S.A., 97, 8392–8396CrossRefGoogle Scholar
, C. R. Woese (2002). On the evolution of cells. Proc. Nat. Acad. Sci. U.S.A., 99, 8742–8747CrossRefGoogle Scholar
, C. R. Woese (2004). A new biology for a new century. Microbiol. Mol. Biol. Rev., 68, 173–186CrossRefGoogle Scholar
, C. R. Woese and , G. E. Fox (1977). Phylogenetic structure of the prokaryotic domain. Proc. Nat. Acad. Sci. U.S.A., 74, 5088–5090CrossRefGoogle Scholar
, J. T.-F. Wong (1975). A co-evolution theory of the genetic code. Proc. Nat. Acad. Sci. U.S.A., 72, 1909–1912CrossRefGoogle Scholar
, J. T.-F. Wong (1991). Origin of genetically encoded protein synthesis: A model based on selection for RNA peptidation. Orig. Life Evol. Biosph., 21, 165–176CrossRefGoogle Scholar
J. T.-F. Wong and H. Xue (2002). Self-perfecting evolution of heteropolymer building blocks and sequences as the basis of life. In Fundamentals of Life (, G. Palyi, , C. Zucchi, and , L. Caglioti, eds.), 473–494. Paris: ElsevierGoogle Scholar
, Y. Xu and , N. Glansdorff (2002). Was our ancestor a hyperthermophilic procaryote?Comp. Biochem. Physiol. Part A, 133, 677–688CrossRefGoogle Scholar
, Y. Yamagata, , H. Watanabe, , M. Saitoh, and , T. Namba (1991). Volcanic production of polyphosphates and its relevance to prebiotic evolution. Nature, 352, 516–519CrossRefGoogle Scholar
, M. Yarus (2000). RNA–ligand chemistry: A testable source for the genetic code. RNA, 6, 475–484CrossRefGoogle Scholar
, S. Yokoyama, , A. Koyama, , A. Nemoto, , H. Honda, , E. Imai, , K. Hatori, and , K. Matsuno (2003). Amplification of diverse catalytic properties of evolving molecules in a simulated hydrothermal environment. Orig. Life Evol. Biosph., 33, 589–595CrossRefGoogle Scholar
, W. Zillig (1991). Comparative biochemistry of Archaea and Bacteria. Curr. Opin. Genet. Dev., 1, 544–551CrossRefGoogle Scholar
, A. Akhmanova, , F. Voncken, , T. van Halen, , A. van Hoek, , B. Boxma, , G. Vogels, , M. Veenhuis, and , H. H. P. Hackstein (1998). A hydrogenosome with a genome. Nature, 396, 527–528CrossRefGoogle Scholar
, A. D. Anbar and , A. H. Knoll (2002). Proterozoic ocean chemistry and evolution: A bioinorganic bridge. Science, 297, 1137–1142CrossRefGoogle Scholar
, J. O. Andersson, , A. M. Sjögren, , L. A. M. Davis, , T. M. Embley, and , A. J. Roger (2003). Phylogenetic analyses of diplomonad genes reveal frequent lateral gene transfers affecting eukaryotes. Curr. Biol., 13, 94–104CrossRefGoogle Scholar
, L. Aravind, , R. L. Tatusov, , Y. I. Wolf, , D. R. Walker, and , E. V. Koonin (1998). Evidence for massive gene exchange between archaeal and bacterial hyperthermophiles. Trends Genet., 14, 442–444CrossRefGoogle Scholar
, G. L. Arnold, , A. D. Anbar, , J. Barling, and , T. W. Lyons (2004). Molybdenum isotope evidence for widespread anoxia in mid-proterozoic oceans. Science, 304, 87–90CrossRefGoogle Scholar
G. Arrhenius, B. Gedulin, and S. Mojzsis (1993). Phosphate in models for chemical evolution. In Chemical Evolution and Origin of Life (, C. Ponnamperuma and , J. Chela-Flores, eds.), 25–50. Hampton VA: A. Deepak PublishersGoogle Scholar
, M. Balter (2000). Evolution on life's fringes. Science, 289, 1866–1867CrossRefGoogle Scholar
M. Baltscheffsky and H. Baltscheffsky (1992). Inorganic pyrophosphate and inorganic pyrophosphatases. In Molecular Mechanisms in Bioenergetics (, L. Ernster, ed.), 331–348. Amsterdam: ElsevierGoogle Scholar
, A. Bar-Nun, , E. Kochavi, and , S. Bar-Nun (1994). Assemblies of free amino acids as possible prebiotic catalysts. J. Mol. Evol., 39, 116–122Google Scholar
, A. D. Baughn and , M. H. Malamy (2004). The strict anaerobe Bacteroides fragilis grows in and benefits from nanomolar concentrations of oxygen. Nature, 427, 441–444CrossRefGoogle Scholar
, A. Bekker, , H. D. Holland, , P.-L. Wang, , D. Rumble III, , H. J. Stein, , J. L. Hannah, , L. L. Coetzee, and , N. J. Beukes (2004). Dating the rise of atmospheric oxygen. Nature, 427, 117–120CrossRefGoogle Scholar
, M. P. Bernstein, , J. P. Dworkin, , S. A. Sandford, , G. W. Cooper, and , L. J. Allamandola (2002). Racemic amino acids from the ultraviolet photolysis of interstellar ice analogues. Nature, 416, 401–403CrossRefGoogle Scholar
, O. Botta and , J. L. Bada (2002). Extraterrestrial organic compounds in meteorites. Surv. Geophys., 23, 411–467CrossRefGoogle Scholar
, O. Botta, , D. P. Glavin, , G. Kminek, and , J. L. Bada (2002). Relative amino acid concentrations as a signature for parent body processes of carbonaceous chondrites. Orig. Life Evol. Biosph., 32, 143–164CrossRefGoogle Scholar
, Y. Boucher, , M. Kamekura, and , W. F. Doolittle (2004). Origins and evolution of isoprenoid lipid biosynthesis in archaea. Mol. Microbiol., 52 (No. 2), 515–527CrossRefGoogle Scholar
A. Brack (2003). La chimie de l'origine de la vie. In Les Traces du Vivant (, M. Gargaud, , D. Despois, , J.-P. Parisot, and , J. Reisse, eds.), 61–81. Pessac: Presses Universitaires de BordeauxGoogle Scholar
, M. D. Brasier, , O. R. Green, , A. P. Jephcoat, , A. K. Kleppe, , M. J. Van Kranendonk, , J. F. Lindsay, , A. Steele, and , N. V. Grassineau (2002). Questioning the evidence for Earth's oldest fossils. Nature, 416, 76–81CrossRefGoogle Scholar
, J. J. Brocks, , G. A. Logan, , R. Buick, and , R. E. Summons (1999). Archaean molecular fossils and the early rise of eukaryotes. Science, 285, 1033–1036CrossRefGoogle Scholar
, P. Brown, , T. Sutikna, , M. J. Morwood, , R. P. Soejono, , E. Wayhu Saptomo, and , R. Awe Due (2004). A new small-bodied hominin from the late Pleistocene of Flores, Indonesia. Nature, 431, 1055–1061CrossRefGoogle Scholar
, R. Cammack (1983). Evolution and diversity in the iron–sulfur proteins. Chem. Scr., 21, 87–95Google Scholar
, D. E. Cane (1997). Polyketide and nonribosomal polypeptide biosynthesis. Chem. Rev., 97, 2463–2706. (includes 13 papers on the topic.)CrossRefGoogle Scholar
, D. E. Canfield (1998). A new model for Proterozoic ocean chemistry. Nature, 396, 450–453CrossRefGoogle Scholar
, L. H. Caporale (2003). Foresight in genome evolution. Am. Sci., 91, 234–241CrossRefGoogle Scholar
, D. Caramelli, , C. Lalueza-Fox, , C. Vernes, , M. Lari, , A. Casoli, , F. Mallegni, , B. Chiarelli, , I. Dupanloup, , J. Bertranpetit, , G. Barbujani, and , G. Bertorelle (2003). Evidence for a genetic discontinuity between Neandertals and 24,000-year-old anatomically modern Europeans. Proc. Nat. Acad. Sci. U.S.A., 100, 6593–6597CrossRefGoogle Scholar
, S. B. Carroll (2003). Genetics and the making of Homo sapiens. Nature, 422, 849–857CrossRefGoogle Scholar
, J. Castresana and , M. Saraste (1995). Evolution of energetic metabolism: The respiration early hypothesis. Trends Biol. Sci., 20, 443–448CrossRefGoogle Scholar
, T. Cavalier-Smith (1975). The origin of nuclei and eukaryotic cells. Nature, 256, 463–468CrossRefGoogle Scholar
, T. Cavalier-Smith (2002). The phagotrophic origin of eukaryotes and phylogenetic classification of Protozoa. Int. J. Syst. Evol. Microbiol., 52, 297–354CrossRefGoogle Scholar
, T. R. Cech (1986). RNA as an enzyme. Sci. Am., 255 (No. 5), 64–75CrossRefGoogle Scholar
, A. Chakrabarti, , R. R. Breaker, , G. F. Joyce, and , D. W. Deamer (1994). Production of RNA by a polymerase protein encapsulated within phospholipid vesicles. J. Mol. Evol., 39, 555–559CrossRefGoogle Scholar
, S. Conway Morris (1998). The Crucible of Creation. Oxford: Oxford University PressGoogle Scholar
, S. Conway Morris (2003). Life's Solution. Cambridge: Cambridge University PressGoogle Scholar
, J. R. Cronin and , S. Pizzarello (1997). Enantiomeric excesses in meteoritic amino acids. Science, 275, 951–955CrossRefGoogle Scholar
, C. Cunchillos and , G. Lecointre (2002). Early steps of metabolism evolution inferred by cladistic analysis of amino acid catabolic pathways. C. R. Biol., 325, 119–129CrossRefGoogle Scholar
, C. Cunchillos and , G. Lecointre (2005). Integrating the universal metabolism into a phylogenetic analysis. Mol. Biol. Evol., 22, 1–11Google Scholar
D. W. Deamer (1998). Membrane compartments in prebiotic evolution. In The Molecular Origins of Life (, A. Brack, ed.), 189–205. Cambridge: Cambridge University PressGoogle Scholar
C. de Duve (1963). The lysosome concept. In Ciba Foundation Symposium on Lysosomes (, A. V. S. de Reuck and , M. P. Cameron, eds.), 1–31. London: ChurchillGoogle Scholar
, C. de Duve (1969). Evolution of the peroxisome. Ann. N.Y. Acad. Sci., 168, 369–381CrossRefGoogle Scholar
, C. de Duve (1984). A Guided Tour of the Living Cell. New York: Scientific American BooksGoogle Scholar
, C. de Duve (1988). The second genetic code. Nature, 333, 117–118CrossRefGoogle Scholar
, C. de Duve (1991). Blueprint for a Cell. Burlington, NC: Neil Patterson Publishers, Carolina Biological Supply CompanyGoogle Scholar
, C. de Duve (1993). The RNA world: Before and after?Gene, 135, 29–31CrossRefGoogle Scholar
, C. de Duve (1995). Vital Dust. New York: BasicBooksGoogle Scholar
, C. de Duve (1996). The birth of complex cells. Sci. Am., 274 (No. 4), 38–45Google Scholar
C. de Duve (1998). Clues from present-day biology: The thioester world. In The Molecular Origins of Life (, A. Brack, ed.), 219–236. Cambridge: Cambridge University PressGoogle Scholar
C. de Duve (2001). The origin of life: Energy. In Frontiers of Life. (, D. Baltimore, , R. Dulbecco, , F. Jacob, and , R. Levi-Montalcini, eds.), Vol. I, 153–168. San Diego, CA: Academic PressGoogle Scholar
, C. de Duve (2002). Life Evolving. New York: Oxford University PressGoogle Scholar
, C. de Duve (2003). A research proposal on the origin of life. Orig. Life Evol. Biosph., 33, 1–16CrossRefGoogle Scholar
, C. de Duve (2005). The onset of selection. Nature, 433, 581–582CrossRefGoogle Scholar
, C. de Duve and , R. Wattiaux (1966). Functions of lysosomes. Annu. Rev. Physiol., 28, 435–492CrossRefGoogle Scholar
, F. M. Devienne, , C. Barnabé, , M. Couderc, and , G. Ourisson (1998). Synthesis of biological compounds in quasi-interstellar conditions. C. R. Acad. Sci. Paris, Sér. IIc, 1, 435–439Google Scholar
, F. M. Devienne, , C. Barnabé, and , G. Ourisson (2002). Synthesis of further biological compounds in interstellar-like conditions. C. R. Acad. Sci. Paris, Chimie/Chemistry, 5, 651–653Google Scholar
, R. E. Dickerson and , I. Geis (1969). The Structure and Action of Proteins. Menlo Park, CA: Benjamin/Cummings Publishing CompanyGoogle Scholar
, M. Di Giulio (2003). The early phases of genetic code origin: Conjectures on the evolution of coded catalysis. Orig. Life Evol. Biosph., 33, 479–489CrossRefGoogle Scholar
, W. F. Doolittle (1999). Phylogenetic classification and the universal tree. Science, 284, 2124–2128CrossRefGoogle Scholar
, W. F. Doolittle (2000). The nature of the universal ancestor and the evolution of the proteome. Curr. Opin. Struct. Biol., 10, 355–358CrossRefGoogle Scholar
, S. D. Dyall, , M. T. Brown, and , P. J. Johnson (2004a). Ancient invasions: From endosymbionts to organelles. Science, 304, 253–257CrossRefGoogle Scholar
, S. D. Dyall, , W. Yan, , M. G. Delgadillo-Correa, , A. Lunceford, , J. A. Loo, , C. F. Clarke, and , P. J. Johnson (2004b). Non-mitochondrial complex I proteins in a Trichomonas hydrogenosomal oxidoreductase complex. Nature, 431, 1103–1107CrossRefGoogle Scholar
, P. Ehrenfreund, , W. Irvine, , L. Becker, , J. Blank, , J. R. Brucato, , L. Colangeli, , S. Derenne, , D. Despois, , A. Dutrey, , H. Fraaije, , A. Lazcano, , T. Owen, and , F. Robert, an International Space Science Institute ISSI Team (2002). Astrophysical and astrochemical insights into the origin of life. Rep. Prog. Phys., 65, 1427–1487CrossRefGoogle Scholar
, M. Eigen and , P. Schuster (1977). The hypercycle: A principle of self-organization. Part A: Emergence of the hypercycle. Naturwissenschaften, 64, 541–565CrossRefGoogle Scholar
, M. Eigen and , R. Winkler-Oswatitsch (1981). Transfer-RNA, an early gene. Naturwissenschaften, 68, 282–292CrossRefGoogle Scholar
, A. Eschenmoser (1999). Chemical etiology of nucleic acid structure. Science, 284, 2118–2124CrossRefGoogle Scholar
, P. Forterre (1995). Thermoreduction, a hypothesis for the origin of prokaryotes. C. R. Acad. Sci. III, 318, 415–422Google Scholar
, P. Forterre (1999). Where is the root of the universal tree of life?BioEssays, 21, 871–8793.0.CO;2-Q>CrossRefGoogle Scholar
, P. Forterre, , C. Bouthier de la Tour, , H. Philippe, and , M. Duguet (2000). Reverse gyrase from hyperthermophiles: Probable transfer of a thermoadaptation trait from Archaea to Bacteria. Trends Genet., 16, 152–154CrossRefGoogle Scholar
, S. W. Fox (1988). The Emergence of Life. New York: BasicBooksGoogle Scholar
, S. J. Freeland, , T. Wu, and , N. Keulmann (2003). The case for an error minimizing standard genetic code. Orig. Life Evol. Biosph., 33, 457–477CrossRefGoogle Scholar
, N. Fujii (2002). d-Amino acids in living higher organisms. Orig. Life Evol. Biosph., 32, 103–127CrossRefGoogle Scholar
, H. Furnes, , N. R. Banerjee, , K. Muehlenbachs, , H. Staudigel, and , M. de Wit (2004). Early life recorded in archean pillow lavas. Science, 304, 578–581CrossRefGoogle Scholar
, N. Galtier, , N. Tourasse, and , M. Gouy (1999). A nonhyperthermophilic common ancestor to extant life forms. Science, 283, 220–221CrossRefGoogle Scholar
, J. M. Garcia-Ruiz, , S. T. Hyde, , A. M. Carnerup, , A. G. Christy, , M. J. Van Kranendonk, and , N. J. Welham (2003). Self-assembled silica–carbonate structures and detection of ancient microfossils. Science, 302, 1194–1197CrossRefGoogle Scholar
B. Gedulin and G. Arrhenius (1994). Sources and geochemical evolution of RNA precursor molecules – the role of phosphate. In Early Life on Earth, Nobel Symposium 84 (, S. Bengtson, ed.), 91–110. New York: Columbia University PressGoogle Scholar
, W. Gilbert (1986). The RNA world. Nature, 319, 618CrossRefGoogle Scholar
, W. Gilbert, , M. Marchionni, and , G. Mcknight (1986). On the antiquity of introns. Cell, 46, 151–154CrossRefGoogle Scholar
, N. Glansdorff (2000). About the last common ancestor, the universal life-tree, and lateral gene transfer: A reappraisal. Mol. Microbiol., 38 (No. 2), 177–185CrossRefGoogle Scholar
, M. Gogarten-Boekels, , E. Hilario, and , J. P. Gogarten (1995). The effects of heavy meteorite bombardment on the early evolution – the emergence of the three domains of life. Orig. Life Evol. Biosph., 25, 251–264CrossRefGoogle Scholar
, S. Gribaldo and , H. Philippe (2002). Ancient phylogenetic relationships. Theor. Pop. Biol., 61, 391–408CrossRefGoogle Scholar
, R. S. Gupta, , K. Aitken, , M. Falah, and , B. Singh (1994). Cloning of Giardia lamblia heat shock protein HSP70 homologs: Implications regarding origin of eukaryotic cells and endoplasmic reticulum. Proc. Nat. Acad. Sci. U.S.A., 91, 2895–2899CrossRefGoogle Scholar
, J. H. P. Hackstein, , A. Akhmanova, , F. Voncken, , A. van Hoek, , T. van Alen, , B. Boxma, , S. Y. Moon-van der Staay, , G. van der Staay, , J. Leunissen, , M. Huynen, , J. Rosenberg, and , M. Veenhuis (2001). Hydrogenosomes: Convergent adaptations of mitochondria to anaerobic environments. Zoology, 104, 290–302CrossRefGoogle Scholar
, M. M. Hanczyc, , S. M. Fujikawa, and , J. W. Szostak (2003). Experimental models of primitive cellular compartments: Encapsulation, growth, and division. Science, 302, 618–622CrossRefGoogle Scholar
, H. Hartman (1984). The origin of the eukaryotic cell. Speculations Sci. Technol., 7 (No. 2), 77–81Google Scholar
, H. Hartman and , A. Fedorov (2002). The origin of the eukaryotic cell: A genomic investigation. Proc. Nat. Acad. Sci. U.S.A., 99, 1420–1425CrossRefGoogle Scholar
, T. Horiike, , K. Hamada, , S. Kanaya, and , T. Shinozawa (2001). Origin of eukaryotic cell nuclei by symbiosis of Archaea in Bacteria is revealed by homology-hit analysis. Nature Cell Biol., 3, 210–214CrossRefGoogle Scholar
, D. S. Horner, , P. G. Foster, and , T. M. Embley (2000). Iron hydrogenases and the evolution of anaerobic eukaryotes. Mol. Biol. Evol., 17 (No. 11), 1695–1709CrossRefGoogle Scholar
, C. Huber and , G. Wächtershäuser (1998). Peptides by activation of amino acids by CO on (Ni, Fe)S surfaces: Implications for the origin of life. Science, 281, 670–672CrossRefGoogle Scholar
, E. Imai, , H. Honda, , K. Hatori, , A. Brack, and , K. Matsuno (1999). Elongation of oligopeptides in a simulated submarine hydrothermal system. Science, 283, 831–833CrossRefGoogle Scholar
, W. K. Johnston, , P. J. Unrau, , M. S. Lawrence, , M. E. Glasner, and , D. P. Bartel (2001). RNA-catalyzed RNA polymerization: Accurate and general RNA-templated primer extension. Science, 292, 1319–1325CrossRefGoogle Scholar
, A. Jorissen and , C. Cerf (2002). Asymmetric photoreactions as the origin of biomolecular homochirality: A critical review. Orig. Life Evol. Biosph., 32, 129–142CrossRefGoogle Scholar
, G. F. Joyce (2002). The antiquity of RNA-based evolution. Nature, 418, 214–221CrossRefGoogle Scholar
G. F. Joyce and L. E. Orgel (1993). Prospects for the understanding of the origin of the RNA world. In The RNA World (, R. F. Gesteland and , J. F. Atkins, eds.), 1–25. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory PressGoogle Scholar
, O. Kandler (1994a). Cell wall biochemistry in Archaea and its phylogenetic implications. J. Biol. Phys., 20, 165–169Google Scholar
O. Kandler (1994b). The early diversification of life. In Early Life on Earth, Nobel Symposium 84 (, S. Bengtson, ed.), 152–160. New York: Columbia University PressGoogle Scholar
, K. Kashefi and , D. R. Lovley (2003). Extending the upper temperature limit for life. Science, 301, 934CrossRefGoogle Scholar
M. Kates (1992). Archaebacterial lipids: Structure, biosynthesis and function. In The Archaebacteria: Biochemistry and Biotechnology (, M. J. Danson, , D. W. Hough, and , G. G. Lunt, eds.), 51–72. Biochem. Soc. Symp. 58. London: Portland PressGoogle Scholar
, A. D. Keefe and , S. L. Miller (1995). Are polyphosphates or phosphate esters prebiotic reagents?J. Mol. Evol., 41, 693–702Google Scholar
, A. D. Keefe, , G. L. Newton, and , S. L. Miller (1995). A possible prebiotic synthesis of pantetheine, a precursor of coenzyme A. Nature, 373, 683–685CrossRefGoogle Scholar
, A. H. Knoll (2003). Life on a Young Planet. Princeton, NJ: Princeton University PressGoogle Scholar
, V. Kolb, , S. Zhang, , Y. Xu, and , G. Arrhenius (1997). Mineral-induced phosphorylation of glycolate ion – a metaphor in chemical evolution. Orig. Life Evol. Biosph., 27, 485–503CrossRefGoogle Scholar
, E. V. Koonin (2003). Comparative genomics, minimal gene-sets and the last universal common ancestor. Nature Rev. Microbiol., 1, 127–136CrossRefGoogle Scholar
, A. Kornberg, , N. N. Rao, and , D. Ault-Riché (1999). Inorganic polyphosphate: A molecule of many functions. Annu. Rev. Biochem., 68, 89–125CrossRefGoogle Scholar
, I. S. Kulaev (1979). The Biochemistry of Inorganic Polyphosphates. New York: WileyGoogle Scholar
, J. A. Lake, , R. Jain, and , M. C. Rivera (1999). Mix and match in the tree of life. Science, 283, 2027–2028CrossRefGoogle Scholar
, N. Lane (2002). Oxygen. Oxford: Oxford University PressGoogle Scholar
, A. Lazcano (2003). Just how pregnant is the universe?Science, 299, 347–348CrossRefGoogle Scholar
, L. Leman, , L. Orgel, and , M. Reza Ghadiri (2004). Carbonyl sulfide-mediated prebiotic formation of peptides. Science, 306, 283–286CrossRefGoogle Scholar
, R. Lewin (1996). Patterns in Evolution. New York: Scientific American BooksGoogle Scholar
, M. R. Lindsay, , R. I. Webb, , M. Strous, , M. S. Jetten, , M. K. Butler, , R. J. Forde, and , J. A. Fuerst (2001). Cell compartmentalisation in planctomycetes: Novel types of structural organisation for the bacterial cell. Arch. Microbiol., 175 (No. 6), 413–429CrossRefGoogle Scholar
P. L. Luisi (2002). Some open questions about the origin of life. In Fundamentals of Life (, G. Palyi, , C. Zucchi, and , L. Caglioti, eds.), 289–301. Paris: ElsevierGoogle Scholar
, D. A. Mac Donaill (2003). Why nature chose A, C, G, and U/T: An error-coding perspective of nucleotide alphabet composition. Orig. Life Evol. Biosph., 33, 433–455CrossRefGoogle Scholar
, L. Margulis (1981). Symbiosis in Cell Evolution. San Francisco: W. H. Freeman & CoGoogle Scholar
, L. Margulis (1996). Archaeal–eubacterial mergers in the origin of Eukarya: Phylogenetic classification of life. Proc. Nat. Acad. Sci. U.S.A., 93, 1071–1076CrossRefGoogle Scholar
, L. Margulis and , D. Sagan (1986). Micro-cosmos. New York: Summit BooksGoogle Scholar
, W. Martin (1999). Mosaic bacterial chromosomes: A challenge en route to a tree of genomes. BioEssays, 21, 99–1043.0.CO;2-B>CrossRefGoogle Scholar
, W. Martin and , M. Müller (1998). The hydrogen hypothesis for the first eukaryote. Nature, 392, 37–41CrossRefGoogle Scholar
, W. Martin, , C. Rotte, , M. Hoffmeister, , U. Theissen, , G. Gelius-Dietrich, , S. Ahr, and , K. Henze (2003). Early cell evolution, eukaryotes, anoxia, sulfide, oxygen, fungi first (?), and a tree of genomes revisited. Life, 55 (No. 4–5), 193–204Google Scholar
, W. Martin and , M. J. Russell (2003). On the origins of cells: A hypothesis for the evolutionary transitions from abiotic geochemistry to chemoautotrophic prokaryotes, and from prokaryotes to nucleated cells. Phil. Trans. R. Soc. London B, 358, 59–85CrossRefGoogle Scholar
, S. L. Miller (1953). A production of amino acids under possible primitive earth conditions. Science, 117, 528–529CrossRefGoogle Scholar
, K. Miller (1999). Finding Darwin's God. New York: HarperCollinsGoogle Scholar
, M. Mirazon Lahr and , R. Foley (2004). Human evolution writ small. Nature, 431, 1043–1044CrossRefGoogle Scholar
, P. A. Monnard, , C. L. Apel, , A. Kanavarioti, and , D. W. Deamer (2004). Influence of ionic inorganic solutes on self-assembly and polymerization processes related to early forms of life: Implications for a prebiotic aqueous medium. Astrobiology, 2 (No. 2), 139–152Google Scholar
, D. Moreira and , P. Lopez-Garcia (1998). Symbiosis between methanogenic archaea and δ-proteobacteria as the origin of eukaryotes: The syntrophic hypothesis. J. Mol. Evol., 47, 517–530CrossRefGoogle Scholar
, H. J. Morowitz (1999). A theory of biochemical organization, metabolic pathways, and evolution. Complexity, 4 (No. 6), 39–533.0.CO;2-2>CrossRefGoogle Scholar
, M. J. Morwood, , R. P. Soejono, , R. G. Roberts, , T. Sutikna, , C. S. M. Turney, , K. E. Westaway, , W. J. Rink, , J.-x. Zhao, , G. D. van den Bergh, , R. Awe Due, , D. R. Hobbs, , M. W. Moore, , M. I. Bird, and , L. K. Fifield (2004). Archaeology and age of a new hominin from Flores in eastern Indonesia. Nature, 431, 1087–1091CrossRefGoogle Scholar
, M. Müller (1993). The hydrogenosome. J. Gen. Microbiol., 139, 2879–2889CrossRefGoogle Scholar
, M. Müller and , W. Martin (1999). The genome of Rickettsia prowazekii and some thoughts on the origin of mitochondria and hydrogenosomes. BioEssays, 21, 377–3813.0.CO;2-W>CrossRefGoogle Scholar
, G. M. Munoz Caro, , U. J. Meierhenrich, , W. A. Schutte, , B. Barbier, , A. Arcones Segovia, , H. Rosenbauer, , W. H.-P. Thiemann, , A. Brack, and , J. M. Greenberg (2002). Amino acids from ultraviolet irradiation of interstellar ice analogues. Nature, 416, 403–406CrossRefGoogle Scholar
, K. E. Nelson, , R. A. Clayton, , S. R. Gill, , M. L. Gwinn, , R. J. Dodson, , D. H. Haft, , E. K. Hickey, , J. D. Peterson, , W. C. Nelson, , K. A. Ketchum, , L. McDonald, , T. R. Utterback, , J. A. Malek, , K. D. Linher, , M. M. Garrett, , A. M. Stewart, , M. D. Cotton, , M. S. Pratt, , C. A. Phillips, , D. Richardson, , J. Heidelberg, , G. G. Sutton, , R. D. Fleischmann, , J. A. Eisen, , O. White, , S. L. Salzberg, , H. O. Smith, , J. C. Venter, and , C. M. Fraser (1999). Evidence for lateral gene transfer between Archaea and Bacteria from genome sequence of Thermotoga maritima. Nature, 399, 323–329CrossRefGoogle Scholar
, E. Nevo (1999). Mosaic Evolution of Subterranean Mammals. Oxford: Oxford University PressGoogle Scholar
, E. Nogales, , K. H. Downing, , L. A. Amos, and , J. Löwe (1998). Tubulin and FtsZ form a distinct family of GTPases. Nature Struct. Biol., 5, 451–458CrossRefGoogle Scholar
, H. Ochman, , J. G. Lawrence, and , E. A. Groisman (2000). Lateral gene transfer and the nature of bacterial evolution. Nature, 405, 299–304CrossRefGoogle Scholar
, H. Ogasawara, , A. Yoshida, , E. Imai, , H. Honda, , K. Hatori, and , K. Matsuno (2000). Synthesizing oligomers from monomeric nucleotides in simulated hydrothermal environments. Orig. Life Evol. Biosph., 30, 519–526CrossRefGoogle Scholar
, Y. Ogata, , E. Imai, , H. Honda, , K. Hatori, and , K. Matsuno (2000). Hydrothermal circulation of sea water through hot vents and contribution of interface chemistry to prebiotic synthesis. Orig. Life Evol. Biosph., 30, 527–537CrossRefGoogle Scholar
, L. E. Orgel (2003). Some consequences of the RNA world hypothesis. Orig. Life Evol. Biosph., 33, 211–218CrossRefGoogle Scholar
S. Osawa (1995). Evolution of the Genetic Code. Oxford University Press
, G. Ourisson and , T. Nakatani (1994). The terpenoid theory of the origin of cellular life: The evolution of terpenoids to cholesterol. Chemistry and Biology, 1, 11–23CrossRefGoogle Scholar
, K. Ozawa, , A. Nemoto, , E. Imai, , H. Honda, , K. Hatori, and , K. Matsuno (2004). Phosphorylation of nucleotide molecules in hydrothermal environments. Orig. Life Evol. Biosph., 34, 465–471CrossRefGoogle Scholar
, J. D. Palmer (2003). The symbiotic birth and spread of plastids: How many times and whodunit?J. Phycol., 39, 4–11CrossRefGoogle Scholar
, S. Pitsch, , A. Eschenmoser, , B. Gedulin, , S. Hui, and , G. Arrhenius (1995). Mineral induced formation of sugar phosphates. Orig. Life Evol. Biosph., 25, 294–334Google Scholar
, S. Pizzarello and , A. L. Weber (2004). Prebiotic amino acids as asymmetric catalysts. Science, 303, 1151CrossRefGoogle Scholar
, A. Poole, , D. Jeffares, and , D. Penny (1999). Early evolution: prokaryotes, the new kids on the block. Bioessays, 21, 880–8893.0.CO;2-P>CrossRefGoogle Scholar
, D. Prangishvili (2003). Evolutionary insights from studies on viruses of hyperthermophilic archaea. Res. Microbiol., 154, 289–294CrossRefGoogle Scholar
, B. P. Prieur (2001). Étude de l'activité prébiotique potentielle de l'acide borique. C. R. Acad. Sci. Paris, Chimie/Chemistry, 4, 1–4Google Scholar
, A. Ricardo, , M. A. Carrigan, , A. N. Olcott, and , S. A. Benner (2004). Borate minerals stabilize ribose. Science, 303, 196CrossRefGoogle Scholar
, M. C. Rivera and , J. A. Lake (2004). The ring of life provides evidence for a genome fusion origin of eukaryotes. Nature, 431, 152–155CrossRefGoogle Scholar
, M. Rohmer (1999). The discovery of a mevalonate-independent pathway for isoprenoid biosynthesis in bacteria, algae, and higher plants. Nat. Prod. Rep., 16, 565–574CrossRefGoogle Scholar
, M. Rohmer, , C. Grosdemange-Billiard, , M. Seemann, and , D. Tritsch (2004). Isoprenoid biosynthesis as a novel target for antibacterial and antiparasitic drugs. Curr. Opin. Investig. Drugs, 5 (No. 2), 154–162Google Scholar
, L. Sagan (1967). On the origin of mitosing cells. J. Theor. Biol., 14, 225–274CrossRefGoogle Scholar
, J. G. Schmidt, , P. E. Nielsen, and , L. E. Orgel (1997). Enantiomeric cross-inhibition in the synthesis of oligonucleotides on a nonchiral template. J. Am. Chem. Soc., 119, 1494–1495CrossRefGoogle Scholar
, J. W. Schopf (1999). Cradle of Life. Princeton, NJ: Princeton University PressGoogle Scholar
A. W. Schwartz (1998). Origins of the RNA world. In The Molecular Origins of Life (, A. Brack, ed.), 237–254. Cambridge: Cambridge University PressGoogle Scholar
, M. Shimizu (1982). Molecular basis for the genetic code. J. Mol. Evol., 18, 297–303CrossRefGoogle Scholar
, A. Shimoyama and , R. Ogasawara (2002). Peptides and diketopiperazines in the Yamato-791198 and Murchison carbonaceous chondrites. Orig. Life Evol. Biosph., 32, 165–179CrossRefGoogle Scholar
, N. H. Sleep, , K. J. Zahnle, , J. F. Kasting, and , H. J. Morowitz (1989). Annihilation of ecosystems by large asteroid impacts on the early Earth. Nature, 342, 139–142CrossRefGoogle Scholar
M. L. Sogin, J. D. Silberman, G. Hinkle, and H. G. Morrison (1996). Problems with molecular diversity in the Eukarya. In Evolution of Microbial Life (, D. McL. Roberts, , P. Sharp, , G. Alderson, and , M. A. Collins, eds.), 167–184. Cambridge: Cambridge University PressGoogle Scholar
, S. Spiegelman (1967). An in vitro analysis of a replicating molecule. Am. Sci., 55, 221–264Google Scholar
, R. Y. Stanier (1970). Some aspects of the biology of cells and their possible evolutionary significance. Symp. Soc. Gen. Microbiol., 20, 1–38Google Scholar
, R. Sutak, , P. Dolezal, , H. L. Fiumera, , I. Hrdy, , A. Dancys, , M. Delgadillo-Correa, , P. J. Johnson, , M. Müller, and , J. Tachezy (2004). Mitochondrial-type assembly of Fe–S centers in the hydrogenosomes of the amitochondriate eukaryote Trichomonas vaginalis. Proc. Nat. Acad. Sci. U.S.A., 101, 10368–10373CrossRefGoogle Scholar
, E. Szathmary (2002). The gospel of inevitability. Nature, 419, 779–780CrossRefGoogle Scholar
, J. W. Szostak, , D. P. Bartel, and , P. L. Luisi (2001). Synthesizing life. Nature, 409, 387–390CrossRefGoogle Scholar
, K. Tamura and , P. Schimmel (2004). Chiral-selective aminoacylation of an RNA minihelix. Science, 305, 1253CrossRefGoogle Scholar
, J. Tovar, , G. Leon-Avila, , L. B. Sanchez, , R. Sutak, , J. Tachezy, , M. van der Giezen, , M. Hernandez, , M. Müller, and , J. M. Lucocq (2003). Mitochondrial remnant organelles of Giardia function in iron–sulphur protein maturation. Nature, 426, 172–176CrossRefGoogle Scholar
, F. van den Ent, , L. A. Amos, and , J. Löwe (2001). Prokaryotic origin of the actin cytoskeleton. Nature, 413, 39–44CrossRefGoogle Scholar
, M. A. van Zullen, , A. Lepland, and , G. Arrhenius (2002). Reassessing the evidence for the earliest traces of life. Nature, 418, 627–630CrossRefGoogle Scholar
, T. Vellai and , G. Vida (1999). The origin of eukaryotes: The difference between prokaryotic and eukaryotic cells. Proc. R. Soc. London B, 266, 1571–1577CrossRefGoogle Scholar
, F. Voncken, , B. Boxma, , J. Tjaden, , A. Akhmanova, , M. Huynen, , F. Verbeek, , A. G. M. Tielens, , I. Haferkamp, , H. E. Neuhaus, , G. Vogels, , M. Veenhuis, and , J. H. P. Hackstein (2002). Multiple origins of hydrogenosomes: Functional and phylogenetic evidence from the ADP/ATP carrier of the anaerobic chytrid Neocallimastix sp. Mol. Microbiol., 44 (No. 6), 1441–1454CrossRefGoogle Scholar
, C. D. von Dohlen, , S. Kohler, , S. T. Alsop, and , W. R. McManus (2001). Mealybug beta-proteobacterial endosymbionts contain gamma-proteobacterial symbionts. Nature, 412, 433–436CrossRefGoogle Scholar
G. Wächtershäuser (1998). Origin of life in an iron–sulfur world. In The Molecular Origins of Life (, A. Brack, ed.), 206–218. Cambridge: Cambridge University PressGoogle Scholar
, G. Wächtershäuser (2003). From pre-cells to Eukarya – a tale of two lipids. Mol. Microbiol., 47 (No. 1), 13–22Google Scholar
, C. T. Walsh (2004). Polyketide and nonribosomal peptide antibiotics: Modularity and versatility. Science, 303, 1805–1810CrossRefGoogle Scholar
, J. Washington (2000). The possible role of volcanic aquifers in prebiologic genesis of organic compounds and RNA. Orig. Life Evol. Biosph., 30, 53–79CrossRefGoogle Scholar
, A. L. Weber (2001). The sugar model: Catalysis by amines and amino acid products. Orig. Life Evol. Biosph., 31, 71–86CrossRefGoogle Scholar
, F. H. Westheimer (1987). Why nature chose phosphates. Science, 235, 1173–1178CrossRefGoogle Scholar
, T. D. White, , B. Asfaw, , D. DeGusta, , H. Gilbert, , G. D. Richards, , G. Suwa, and , F. C. Howell (2003). Pleistocene Homo sapiens from Middle Awash, Ethiopea. Nature, 423, 742–747CrossRefGoogle Scholar
, J. Whitfield (2004). Born in a watery commune. Nature, 427, 674–676CrossRefGoogle Scholar
, C. R. Woese (1987). Bacterial evolution. Microbiol. Rev., 51, 221–271Google Scholar
, C. R. Woese (1998). The universal ancestor. Proc. Nat. Acad. Sci. U.S.A., 95, 6854–6859CrossRefGoogle Scholar
, C. R. Woese (2000). Interpreting the universal phylogenetic tree. Proc. Nat. Acad. Sci. U.S.A., 97, 8392–8396CrossRefGoogle Scholar
, C. R. Woese (2002). On the evolution of cells. Proc. Nat. Acad. Sci. U.S.A., 99, 8742–8747CrossRefGoogle Scholar
, C. R. Woese (2004). A new biology for a new century. Microbiol. Mol. Biol. Rev., 68, 173–186CrossRefGoogle Scholar
, C. R. Woese and , G. E. Fox (1977). Phylogenetic structure of the prokaryotic domain. Proc. Nat. Acad. Sci. U.S.A., 74, 5088–5090CrossRefGoogle Scholar
, J. T.-F. Wong (1975). A co-evolution theory of the genetic code. Proc. Nat. Acad. Sci. U.S.A., 72, 1909–1912CrossRefGoogle Scholar
, J. T.-F. Wong (1991). Origin of genetically encoded protein synthesis: A model based on selection for RNA peptidation. Orig. Life Evol. Biosph., 21, 165–176CrossRefGoogle Scholar
J. T.-F. Wong and H. Xue (2002). Self-perfecting evolution of heteropolymer building blocks and sequences as the basis of life. In Fundamentals of Life (, G. Palyi, , C. Zucchi, and , L. Caglioti, eds.), 473–494. Paris: ElsevierGoogle Scholar
, Y. Xu and , N. Glansdorff (2002). Was our ancestor a hyperthermophilic procaryote?Comp. Biochem. Physiol. Part A, 133, 677–688CrossRefGoogle Scholar
, Y. Yamagata, , H. Watanabe, , M. Saitoh, and , T. Namba (1991). Volcanic production of polyphosphates and its relevance to prebiotic evolution. Nature, 352, 516–519CrossRefGoogle Scholar
, M. Yarus (2000). RNA–ligand chemistry: A testable source for the genetic code. RNA, 6, 475–484CrossRefGoogle Scholar
, S. Yokoyama, , A. Koyama, , A. Nemoto, , H. Honda, , E. Imai, , K. Hatori, and , K. Matsuno (2003). Amplification of diverse catalytic properties of evolving molecules in a simulated hydrothermal environment. Orig. Life Evol. Biosph., 33, 589–595CrossRefGoogle Scholar
, W. Zillig (1991). Comparative biochemistry of Archaea and Bacteria. Curr. Opin. Genet. Dev., 1, 544–551CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure [email protected] is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • Bibliography
  • Christian de Duve, Rockefeller University, New York
  • Book: Singularities
  • Online publication: 18 January 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511614736.025
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • Bibliography
  • Christian de Duve, Rockefeller University, New York
  • Book: Singularities
  • Online publication: 18 January 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511614736.025
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • Bibliography
  • Christian de Duve, Rockefeller University, New York
  • Book: Singularities
  • Online publication: 18 January 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511614736.025
Available formats
×